Harmonic number








The harmonic number Hn,1{displaystyle H_{n,1}}H_{{n,1}} with n=⌊x⌋{displaystyle n=lfloor xrfloor }{displaystyle n=lfloor xrfloor } (red line) with its asymptotic limit γ+ln⁡(x){displaystyle gamma +ln(x)}gamma +ln(x) (blue line) where γ{displaystyle gamma }gamma is the Euler–Mascheroni constant.


In mathematics, the n-th harmonic number is the sum of the reciprocals of the first n natural numbers:


Hn=1+12+13+⋯+1n=∑k=1n1k.{displaystyle H_{n}=1+{frac {1}{2}}+{frac {1}{3}}+cdots +{frac {1}{n}}=sum _{k=1}^{n}{frac {1}{k}}.}H_{n}=1+{frac {1}{2}}+{frac {1}{3}}+cdots +{frac {1}{n}}=sum _{k=1}^{n}{frac {1}{k}}.

Harmonic numbers are related to the harmonic mean in that the n-th harmonic number is also n times the reciprocal of the harmonic mean of the first n positive integers.


Harmonic numbers have been studied since antiquity and are important in various branches of number theory. They are sometimes loosely termed harmonic series, are closely related to the Riemann zeta function, and appear in the expressions of various special functions.


The harmonic numbers roughly approximate the natural logarithm function[1]:143 and thus the associated harmonic series grows without limit, albeit slowly. In 1737, Leonhard Euler used the divergence of the harmonic series to provide a new proof of the infinity of prime numbers. His work was extended into the complex plane by Bernhard Riemann in 1859, leading directly to the celebrated Riemann hypothesis about the distribution of prime numbers.


When the value of a large quantity of items has a Zipf's law distribution, the total value of the n most-valuable items is proportional to the n-th harmonic number. This leads to a variety of surprising conclusions regarding the long tail and the theory of network value.


Bertrand's postulate entails that, except for the case n = 1, the harmonic numbers are never integers.[2]









































































































































































































































































































Contents






  • 1 Identities involving harmonic numbers


    • 1.1 Identities involving π




  • 2 Calculation


  • 3 Generating functions


  • 4 Arithmetic properties


  • 5 Applications


  • 6 Generalizations


    • 6.1 Generalized harmonic numbers


    • 6.2 Multiplication formulas


    • 6.3 Hyperharmonic numbers




  • 7 Harmonic numbers for real and complex values


    • 7.1 Alternative, asymptotic formulation


    • 7.2 Special values for fractional arguments


    • 7.3 Relation to the Riemann zeta function




  • 8 See also


  • 9 Notes


  • 10 References


  • 11 External links





Identities involving harmonic numbers


By definition, the harmonic numbers satisfy the recurrence relation


Hn+1=Hn+1n+1.{displaystyle H_{n+1}=H_{n}+{frac {1}{n+1}}.}{displaystyle H_{n+1}=H_{n}+{frac {1}{n+1}}.}

The harmonic numbers are connected to the Stirling numbers of the first kind by the relation


Hn=1n![n+12].{displaystyle H_{n}={frac {1}{n!}}left[{n+1 atop 2}right].}H_{n}={frac {1}{n!}}left[{n+1 atop 2}right].

The functions


fn(x)=xnn!(log⁡x−Hn){displaystyle f_{n}(x)={frac {x^{n}}{n!}}(log x-H_{n})}f_{n}(x)={frac {x^{n}}{n!}}(log x-H_{n})

satisfy the property


fn′(x)=fn−1(x).{displaystyle f_{n}'(x)=f_{n-1}(x).}f_{n}'(x)=f_{n-1}(x).

In particular


f1(x)=x(log⁡x−1){displaystyle f_{1}(x)=x(log x-1)}f_{1}(x)=x(log x-1)

is an integral of the logarithmic function.[citation needed]


The harmonic numbers satisfy the series identity


k=1nHk=(n+1)(Hn+1−1).{displaystyle sum _{k=1}^{n}H_{k}=(n+1)(H_{n+1}-1).}{displaystyle sum _{k=1}^{n}H_{k}=(n+1)(H_{n+1}-1).}


Identities involving π


There are several infinite summations involving harmonic numbers and powers of π:[3]


n=1∞Hnn⋅2n=112π2{displaystyle sum _{n=1}^{infty }{frac {H_{n}}{ncdot 2^{n}}}={frac {1}{12}}pi ^{2}}{displaystyle sum _{n=1}^{infty }{frac {H_{n}}{ncdot 2^{n}}}={frac {1}{12}}pi ^{2}}

n=1∞Hn2(n+1)2=11360π4{displaystyle sum _{n=1}^{infty }{frac {H_{n}^{2}}{(n+1)^{2}}}={frac {11}{360}}pi ^{4}}{displaystyle sum _{n=1}^{infty }{frac {H_{n}^{2}}{(n+1)^{2}}}={frac {11}{360}}pi ^{4}}

n=1∞Hn2n2=17360π4{displaystyle sum _{n=1}^{infty }{frac {H_{n}^{2}}{n^{2}}}={frac {17}{360}}pi ^{4}}{displaystyle sum _{n=1}^{infty }{frac {H_{n}^{2}}{n^{2}}}={frac {17}{360}}pi ^{4}}

n=1∞Hnn3=172π4{displaystyle sum _{n=1}^{infty }{frac {H_{n}}{n^{3}}}={frac {1}{72}}pi ^{4}}{displaystyle sum _{n=1}^{infty }{frac {H_{n}}{n^{3}}}={frac {1}{72}}pi ^{4}}


Calculation


An integral representation given by Euler[4] is


Hn=∫011−xn1−xdx.{displaystyle H_{n}=int _{0}^{1}{frac {1-x^{n}}{1-x}},dx.}H_{n}=int _{0}^{1}{frac {1-x^{n}}{1-x}},dx.

The equality above is straightforward by the simple algebraic identity


1−xn1−x=1+x+⋯+xn−1.{displaystyle {frac {1-x^{n}}{1-x}}=1+x+cdots +x^{n-1}.}{frac {1-x^{n}}{1-x}}=1+x+cdots +x^{n-1}.

Using the substitution x = 1 − u, another expression for Hn is


Hn=∫011−xn1−xdx=∫011−(1−u)nudu=∫01[−k=1n(−1)k(nk)uk−1]du=−k=1n(−1)k(nk)∫01uk−1du=−k=1n(−1)k1k(nk).{displaystyle {begin{aligned}H_{n}&=int _{0}^{1}{frac {1-x^{n}}{1-x}},dx\[6pt]&=int _{0}^{1}{frac {1-(1-u)^{n}}{u}},du\[6pt]&=int _{0}^{1}left[-sum _{k=1}^{n}(-1)^{k}{binom {n}{k}}u^{k-1}right],du\[6pt]&=-sum _{k=1}^{n}(-1)^{k}{binom {n}{k}}int _{0}^{1}u^{k-1},du\[6pt]&=-sum _{k=1}^{n}(-1)^{k}{frac {1}{k}}{binom {n}{k}}.end{aligned}}}{displaystyle {begin{aligned}H_{n}&=int _{0}^{1}{frac {1-x^{n}}{1-x}},dx\[6pt]&=int _{0}^{1}{frac {1-(1-u)^{n}}{u}},du\[6pt]&=int _{0}^{1}left[-sum _{k=1}^{n}(-1)^{k}{binom {n}{k}}u^{k-1}right],du\[6pt]&=-sum _{k=1}^{n}(-1)^{k}{binom {n}{k}}int _{0}^{1}u^{k-1},du\[6pt]&=-sum _{k=1}^{n}(-1)^{k}{frac {1}{k}}{binom {n}{k}}.end{aligned}}}

A closed form expression for Hn is


Hn=Gn−(n+1)⌊Gnn+1⌋{displaystyle {begin{aligned}H_{n}&=G_{n}-(n+1)leftlfloor {frac {G_{n}}{n+1}}rightrfloor end{aligned}}}{displaystyle {begin{aligned}H_{n}&=G_{n}-(n+1)leftlfloor {frac {G_{n}}{n+1}}rightrfloor end{aligned}}}

where


Gn=(n+(n+1)!n)−1(n+1)!{displaystyle {begin{aligned}G_{n}&={frac {{n+(n+1)! choose n}-1}{(n+1)!}}end{aligned}}}{displaystyle {begin{aligned}G_{n}&={frac {{n+(n+1)! choose n}-1}{(n+1)!}}end{aligned}}}



Graph demonstrating a connection between harmonic numbers and the natural logarithm. The harmonic number Hn can be interpreted as a Riemann sum of the integral: 1n+1dxx=ln⁡(n+1).{displaystyle int _{1}^{n+1}{frac {dx}{x}}=ln(n+1).}{displaystyle int _{1}^{n+1}{frac {dx}{x}}=ln(n+1).}


The nth harmonic number is about as large as the natural logarithm of n. The reason is that the sum is approximated by the integral


1n1xdx{displaystyle int _{1}^{n}{frac {1}{x}},dx}{displaystyle int _{1}^{n}{frac {1}{x}},dx}

whose value is ln(n).


The values of the sequence Hn − ln(n) decrease monotonically towards the limit


limn→(Hn−ln⁡n)=γ,{displaystyle lim _{nto infty }left(H_{n}-ln nright)=gamma ,}lim _{nto infty }left(H_{n}-ln nright)=gamma ,

where γ ≈ 0.5772156649 is the Euler–Mascheroni constant. The corresponding asymptotic expansion as n → +∞ is


Hn∼ln⁡n+γ+12n−k=1∞B2k2kn2k=ln⁡n+γ+12n−112n2+1120n4−,{displaystyle H_{n}sim ln {n}+gamma +{frac {1}{2n}}-sum _{k=1}^{infty }{frac {B_{2k}}{2kn^{2k}}}=ln {n}+gamma +{frac {1}{2n}}-{frac {1}{12n^{2}}}+{frac {1}{120n^{4}}}-cdots ,}H_{n}sim ln {n}+gamma +{frac {1}{2n}}-sum _{k=1}^{infty }{frac {B_{2k}}{2kn^{2k}}}=ln {n}+gamma +{frac {1}{2n}}-{frac {1}{12n^{2}}}+{frac {1}{120n^{4}}}-cdots ,

where Bk{displaystyle B_{k}}B_{k} are the Bernoulli numbers.



Generating functions


A generating function for the harmonic numbers is


n=1∞znHn=−ln⁡(1−z)1−z,{displaystyle sum _{n=1}^{infty }z^{n}H_{n}={frac {-ln(1-z)}{1-z}},}{displaystyle sum _{n=1}^{infty }z^{n}H_{n}={frac {-ln(1-z)}{1-z}},}

where ln(z) is the natural logarithm. An exponential generating function is


n=1∞znn!Hn=−ez∑k=1∞1k(−z)kk!=ezEin⁡(z){displaystyle sum _{n=1}^{infty }{frac {z^{n}}{n!}}H_{n}=-e^{z}sum _{k=1}^{infty }{frac {1}{k}}{frac {(-z)^{k}}{k!}}=e^{z}operatorname {Ein} (z)}{displaystyle sum _{n=1}^{infty }{frac {z^{n}}{n!}}H_{n}=-e^{z}sum _{k=1}^{infty }{frac {1}{k}}{frac {(-z)^{k}}{k!}}=e^{z}operatorname {Ein} (z)}

where Ein(z) is the entire exponential integral. Note that


Ein⁡(z)=E1(z)+γ+ln⁡z=Γ(0,z)+γ+ln⁡z{displaystyle operatorname {Ein} (z)=mathrm {E} _{1}(z)+gamma +ln z=Gamma (0,z)+gamma +ln z}{displaystyle operatorname {Ein} (z)=mathrm {E} _{1}(z)+gamma +ln z=Gamma (0,z)+gamma +ln z}

where Γ(0, z) is the incomplete gamma function.



Arithmetic properties


The harmonic numbers have several interesting arithmetic properties. It is well-known that Hn{textstyle H_{n}}{textstyle H_{n}} is an integer if and only if n=1{textstyle n=1}{textstyle n=1}, a result often attributed to Taeisinger.[5] Indeed, using 2-adic valuation, it is not difficult to prove that for n≥2{textstyle ngeq 2}{textstyle ngeq 2} the numerator of Hn{textstyle H_{n}}{textstyle H_{n}} is an odd number while the denominator of Hn{textstyle H_{n}}{textstyle H_{n}} is an even number. More precisely,


Hn=12⌊log2⁡(n)⌋anbn{displaystyle H_{n}={frac {1}{2^{lfloor log _{2}(n)rfloor }}}{frac {a_{n}}{b_{n}}}}{displaystyle H_{n}={frac {1}{2^{lfloor log _{2}(n)rfloor }}}{frac {a_{n}}{b_{n}}}}

with some odd integers an{textstyle a_{n}}{textstyle a_{n}} and bn{textstyle b_{n}}{textstyle b_{n}}.


As a consequence of Wolstenholme's theorem, for any prime number p≥5{displaystyle pgeq 5}pgeq 5 the numerator of Hp−1{displaystyle H_{p-1}}H_{{p-1}}is divisible by p2{textstyle p^{2}}{textstyle p^{2}}. Furthermore, Eisenstein[6] proved that for all odd prime number p{textstyle p}{textstyle p} it holds


H(p−1)/2≡2qp(2)(modp){displaystyle H_{(p-1)/2}equiv -2q_{p}(2){pmod {p}}}{displaystyle H_{(p-1)/2}equiv -2q_{p}(2){pmod {p}}}

where qp(2)=(2p−1−1)/p{textstyle q_{p}(2)=(2^{p-1}-1)/p}{textstyle q_{p}(2)=(2^{p-1}-1)/p} is a Fermat quotient, with the consequence that p{textstyle p}{textstyle p} divides the numerator of H(p−1)/2{displaystyle H_{(p-1)/2}}{displaystyle H_{(p-1)/2}} if and only if p{textstyle p}{textstyle p} is a Wieferich prime.


In 1991, Eswarathasan and Levine[7] defined J(p){displaystyle J(p)}{displaystyle J(p)} as the set of all positive integers n{displaystyle n}n such that the numerator of Hn{displaystyle H_{n}}H_{n} is divisible by a prime number p.{displaystyle p.}p. They proved that


{p−1,p2−p,p2−1}⊆J(p){displaystyle {p-1,p^{2}-p,p^{2}-1}subseteq J(p)}{displaystyle {p-1,p^{2}-p,p^{2}-1}subseteq J(p)}

for all prime numbers p≥5,{displaystyle pgeq 5,}{displaystyle pgeq 5,} and they defined harmonic primes to be the primes p{textstyle p}{textstyle p} such that J(p){displaystyle J(p)}{displaystyle J(p)} has exactly 3 elements.


Eswarathasan and Levine also conjectured that J(p){displaystyle J(p)}{displaystyle J(p)} is a finite set for all primes p,{displaystyle p,}p, and that there are infinitely many harmonic primes. Boyd[8] verified that J(p){displaystyle J(p)}{displaystyle J(p)} is finite for all prime numbers up to p=547{displaystyle p=547}{displaystyle p=547} except 83, 127, and 397; and he gave an heuristic suggesting that the density of the harmonic primes in the set of all primes should be 1/e.{displaystyle 1/e.}1/e. Sanna[9] showed that Jp{displaystyle J_{p}}J_{p} has zero asymptotic density, while Bing-Ling Wu and Yong-Gao Chen[10] proved that the number of elements of Jp{displaystyle J_{p}}J_{p} not exceeding x{displaystyle x}x is at most 3x23+125log⁡p{displaystyle 3x^{{frac {2}{3}}+{frac {1}{25log p}}}}{displaystyle 3x^{{frac {2}{3}}+{frac {1}{25log p}}}}, for all x≥1{displaystyle xgeq 1}x geq 1.



Applications


The harmonic numbers appear in several calculation formulas, such as the digamma function


ψ(n)=Hn−1−γ.{displaystyle psi (n)=H_{n-1}-gamma .}psi (n)=H_{n-1}-gamma .

This relation is also frequently used to define the extension of the harmonic numbers to non-integer n. The harmonic numbers are also frequently used to define γ using the limit introduced earlier:


γ=limn→(Hn−ln⁡(n)),{displaystyle gamma =lim _{nrightarrow infty }{left(H_{n}-ln(n)right)},}{displaystyle gamma =lim _{nrightarrow infty }{left(H_{n}-ln(n)right)},}

although


γ=limn→(Hn−ln⁡(n+12)){displaystyle gamma =lim _{nto infty }{left(H_{n}-ln left(n+{frac {1}{2}}right)right)}}{displaystyle gamma =lim _{nto infty }{left(H_{n}-ln left(n+{frac {1}{2}}right)right)}}

converges more quickly.


In 2002, Jeffrey Lagarias proved[11] that the Riemann hypothesis is equivalent to the statement that


σ(n)≤Hn+(log⁡Hn)eHn,{displaystyle sigma (n)leq H_{n}+(log H_{n})e^{H_{n}},}{displaystyle sigma (n)leq H_{n}+(log H_{n})e^{H_{n}},}

is true for every integer n ≥ 1 with strict inequality if n > 1; here σ(n) denotes the sum of the divisors of n.


The eigenvalues of the nonlocal problem


λφ(x)=∫11φ(x)−φ(y)|x−y|dy{displaystyle lambda varphi (x)=int _{-1}^{1}{frac {varphi (x)-varphi (y)}{|x-y|}},dy}{displaystyle lambda varphi (x)=int _{-1}^{1}{frac {varphi (x)-varphi (y)}{|x-y|}},dy}

are given by λ=2Hn{displaystyle lambda =2H_{n}}lambda =2H_{n}, where by convention, H0=0.{displaystyle H_{0}=0.}H_{0}=0.



Generalizations



Generalized harmonic numbers


The generalized harmonic number of order m of n is given by


Hn,m=∑k=1n1km.{displaystyle H_{n,m}=sum _{k=1}^{n}{frac {1}{k^{m}}}.}H_{n,m}=sum _{k=1}^{n}{frac {1}{k^{m}}}.

The limit as n tends to infinity is finite if m > 1.


Other notations occasionally used include


Hn,m=Hn(m)=Hm(n).{displaystyle H_{n,m}=H_{n}^{(m)}=H_{m}(n).}H_{n,m}=H_{n}^{(m)}=H_{m}(n).

The special case of m = 0 gives Hn,0=n.{displaystyle H_{n,0}=n.}{displaystyle H_{n,0}=n.} The special case of m = 1 is simply called a harmonic number and is frequently written without the m, as


Hn=∑k=1n1k.{displaystyle H_{n}=sum _{k=1}^{n}{frac {1}{k}}.}H_{n}=sum _{k=1}^{n}{frac {1}{k}}.

The smallest natural number k such that kn does not divide the denominator of generalized harmonic number H(k, n) nor the denominator of alternating generalized harmonic number H′(k, n) is, for n=1, 2, ... :


77, 20, 94556602, 42, 444, 20, 104, 42, 76, 20, 77, 110, 3504, 20, 903, 42, 1107, 20, 104, 42, 77, 20, 2948, 110, 136, 20, 76, 42, 903, 20, 77, 42, 268, 20, 7004, 110, 1752, 20, 19203, 42, 77, 20, 104, 42, 76, 20, 370, 110, 1107, 20, ... (sequence A128670 in the OEIS)

In the limit as n → ∞ for m > 1, the generalized harmonic number converges to the Riemann zeta function


limn→Hn,m=ζ(m).{displaystyle lim _{nrightarrow infty }H_{n,m}=zeta (m).}lim _{nrightarrow infty }H_{n,m}=zeta (m).

The related sum k=1nkm{displaystyle sum _{k=1}^{n}k^{m}}sum _{k=1}^{n}k^{m} occurs in the study of Bernoulli numbers; the harmonic numbers also appear in the study of Stirling numbers.


Some integrals of generalized harmonic numbers are


0aHx,2dx=aπ26−Ha{displaystyle int _{0}^{a}H_{x,2},dx=a{frac {pi ^{2}}{6}}-H_{a}}int _{0}^{a}H_{x,2},dx=a{frac {pi ^{2}}{6}}-H_{a}

and



0aHx,3dx=aA−12Ha,2,{displaystyle int _{0}^{a}H_{x,3},dx=aA-{frac {1}{2}}H_{a,2},}int _{0}^{a}H_{x,3},dx=aA-{frac {1}{2}}H_{a,2}, where A is Apéry's constant, i.e. ζ(3).

and


k=1nHk,m=(n+1)Hn,m−Hn,m−1 for m≥0{displaystyle sum _{k=1}^{n}H_{k,m}=(n+1)H_{n,m}-H_{n,m-1}{text{ for }}mgeq 0}{displaystyle sum _{k=1}^{n}H_{k,m}=(n+1)H_{n,m}-H_{n,m-1}{text{ for }}mgeq 0}

Every generalized harmonic number of order m can be written as a function of harmonic of order m-1 using:



Hn,m=∑k=1n−1Hk,m−1k(k+1)+Hn,m−1n{displaystyle H_{n,m}=sum _{k=1}^{n-1}{frac {H_{k,m-1}}{k(k+1)}}+{frac {H_{n,m-1}}{n}}}H_{n,m}=sum _{k=1}^{n-1}{frac {H_{k,m-1}}{k(k+1)}}+{frac {H_{n,m-1}}{n}}   for example: H4,3=H1,21⋅2+H2,22⋅3+H3,23⋅4+H4,24{displaystyle H_{4,3}={frac {H_{1,2}}{1cdot 2}}+{frac {H_{2,2}}{2cdot 3}}+{frac {H_{3,2}}{3cdot 4}}+{frac {H_{4,2}}{4}}}H_{4,3}={frac {H_{1,2}}{1cdot 2}}+{frac {H_{2,2}}{2cdot 3}}+{frac {H_{3,2}}{3cdot 4}}+{frac {H_{4,2}}{4}}

A generating function for the generalized harmonic numbers is


n=1∞znHn,m=Lim⁡(z)1−z,{displaystyle sum _{n=1}^{infty }z^{n}H_{n,m}={frac {operatorname {Li} _{m}(z)}{1-z}},}{displaystyle sum _{n=1}^{infty }z^{n}H_{n,m}={frac {operatorname {Li} _{m}(z)}{1-z}},}

where Lim⁡(z){displaystyle operatorname {Li} _{m}(z)}{displaystyle operatorname {Li} _{m}(z)} is the polylogarithm, and |z| < 1. The generating function given above for m = 1 is a special case of this formula.


A fractional argument for generalized harmonic numbers can be introduced as follows:


For every p,q>0{displaystyle p,q>0}p,q>0 integer, and m>1{displaystyle m>1}m>1 integer or not, we have from polygamma functions:


Hq/p,m=ζ(m)−pm∑k=1∞1(q+pk)m{displaystyle H_{q/p,m}=zeta (m)-p^{m}sum _{k=1}^{infty }{frac {1}{(q+pk)^{m}}}}H_{q/p,m}=zeta (m)-p^{m}sum _{k=1}^{infty }{frac {1}{(q+pk)^{m}}}

where ζ(m){displaystyle zeta (m)}zeta (m) is the Riemann zeta function. The relevant recurrence relation is:


Ha,m=Ha−1,m+1am{displaystyle H_{a,m}=H_{a-1,m}+{frac {1}{a^{m}}}}H_{a,m}=H_{a-1,m}+{frac {1}{a^{m}}}

Some special values are:




H14,2=16−8G−56π2{displaystyle H_{{frac {1}{4}},2}=16-8G-{tfrac {5}{6}}pi ^{2}}H_{{frac {1}{4}},2}=16-8G-{tfrac {5}{6}}pi ^{2} where G is Catalan's constant

H12,2=4−π23{displaystyle H_{{frac {1}{2}},2}=4-{tfrac {pi ^{2}}{3}}}H_{{frac {1}{2}},2}=4-{tfrac {pi ^{2}}{3}}

H34,2=8G+169−56π2{displaystyle H_{{frac {3}{4}},2}=8G+{tfrac {16}{9}}-{tfrac {5}{6}}pi ^{2}}H_{{frac {3}{4}},2}=8G+{tfrac {16}{9}}-{tfrac {5}{6}}pi ^{2}

H14,3=64−27ζ(3)−π3{displaystyle H_{{frac {1}{4}},3}=64-27zeta (3)-pi ^{3}}H_{{frac {1}{4}},3}=64-27zeta (3)-pi ^{3}

H12,3=8−(3){displaystyle H_{{frac {1}{2}},3}=8-6zeta (3)}H_{{frac {1}{2}},3}=8-6zeta (3)

H34,3=(43)3−27ζ(3)+π3{displaystyle H_{{frac {3}{4}},3}={({tfrac {4}{3}})}^{3}-27zeta (3)+pi ^{3}}H_{{frac {3}{4}},3}={({tfrac {4}{3}})}^{3}-27zeta (3)+pi ^{3}



Multiplication formulas


The multiplication theorem applies to harmonic numbers. Using polygamma functions, we obtain



H2x=12(Hx+Hx−12)+ln⁡2{displaystyle H_{2x}={frac {1}{2}}left(H_{x}+H_{x-{frac {1}{2}}}right)+ln 2}{displaystyle H_{2x}={frac {1}{2}}left(H_{x}+H_{x-{frac {1}{2}}}right)+ln 2}

H3x=13(Hx+Hx−13+Hx−23)+ln⁡3,{displaystyle H_{3x}={frac {1}{3}}left(H_{x}+H_{x-{frac {1}{3}}}+H_{x-{frac {2}{3}}}right)+ln 3,}{displaystyle H_{3x}={frac {1}{3}}left(H_{x}+H_{x-{frac {1}{3}}}+H_{x-{frac {2}{3}}}right)+ln 3,}


or, more generally,


Hnx=1n(Hx+Hx−1n+Hx−2n+⋯+Hx−n−1n)+ln⁡n.{displaystyle H_{nx}={frac {1}{n}}left(H_{x}+H_{x-{frac {1}{n}}}+H_{x-{frac {2}{n}}}+cdots +H_{x-{frac {n-1}{n}}}right)+ln n.}{displaystyle H_{nx}={frac {1}{n}}left(H_{x}+H_{x-{frac {1}{n}}}+H_{x-{frac {2}{n}}}+cdots +H_{x-{frac {n-1}{n}}}right)+ln n.}

For generalized harmonic numbers, we have



H2x,2=12(ζ(2)+12(Hx,2+Hx−12,2)){displaystyle H_{2x,2}={frac {1}{2}}left(zeta (2)+{frac {1}{2}}left(H_{x,2}+H_{x-{frac {1}{2}},2}right)right)}H_{2x,2}={frac {1}{2}}left(zeta (2)+{frac {1}{2}}left(H_{x,2}+H_{x-{frac {1}{2}},2}right)right)

H3x,2=19(6ζ(2)+Hx,2+Hx−13,2+Hx−23,2),{displaystyle H_{3x,2}={frac {1}{9}}left(6zeta (2)+H_{x,2}+H_{x-{frac {1}{3}},2}+H_{x-{frac {2}{3}},2}right),}H_{3x,2}={frac {1}{9}}left(6zeta (2)+H_{x,2}+H_{x-{frac {1}{3}},2}+H_{x-{frac {2}{3}},2}right),


where ζ(n){displaystyle zeta (n)}zeta (n) is the Riemann zeta function.



Hyperharmonic numbers


The next generalization was discussed by J. H. Conway and R. K. Guy in their 1995 book The Book of Numbers.[1]:258 Let


Hn(0)=1n.{displaystyle H_{n}^{(0)}={frac {1}{n}}.}H_{n}^{(0)}={frac {1}{n}}.

Then the nth hyperharmonic number of order r (r>0) is defined recursively as


Hn(r)=∑k=1nHk(r−1).{displaystyle H_{n}^{(r)}=sum _{k=1}^{n}H_{k}^{(r-1)}.}H_{n}^{(r)}=sum _{k=1}^{n}H_{k}^{(r-1)}.

In particular, Hn(1){displaystyle H_{n}^{(1)}}{displaystyle H_{n}^{(1)}} is the ordinary harmonic number Hn{displaystyle H_{n}}H_{n}.



Harmonic numbers for real and complex values


The formulae given above,


Hx=∫011−tx1−tdt=−k=1∞(xk)(−1)kk{displaystyle H_{x}=int _{0}^{1}{frac {1-t^{x}}{1-t}},dt=-sum _{k=1}^{infty }{x choose k}{frac {(-1)^{k}}{k}}}{displaystyle H_{x}=int _{0}^{1}{frac {1-t^{x}}{1-t}},dt=-sum _{k=1}^{infty }{x choose k}{frac {(-1)^{k}}{k}}}

are an integral and a series representation for a function that interpolates the harmonic numbers and, via analytic continuation, extends the definition to the complex plane other than the negative integers x. The interpolating function is in fact closely related to the digamma function


Hx=ψ(x+1)+γ,{displaystyle H_{x}=psi (x+1)+gamma ,}{displaystyle H_{x}=psi (x+1)+gamma ,}

where ψ(x) is the digamma, and γ is the Euler-Mascheroni constant. The integration process may be repeated to obtain


Hx,2=−k=1∞(−1)kk(xk)Hk.{displaystyle H_{x,2}=-sum _{k=1}^{infty }{frac {(-1)^{k}}{k}}{x choose k}H_{k}.}{displaystyle H_{x,2}=-sum _{k=1}^{infty }{frac {(-1)^{k}}{k}}{x choose k}H_{k}.}

The Taylor series for the harmonic numbers is


Hx=∑k=2∞(−1)kζ(k)xk−1 for |x|<1{displaystyle H_{x}=sum _{k=2}^{infty }(-1)^{k}zeta (k);x^{k-1}quad {text{ for }}|x|<1}{displaystyle H_{x}=sum _{k=2}^{infty }(-1)^{k}zeta (k);x^{k-1}quad {text{ for }}|x|<1}

which comes from the Taylor series for the digamma function.



Alternative, asymptotic formulation


When seeking to approximate Hx for a complex number x it turns out that it is effective to first compute Hm for some large integer m, then use that to approximate a value for Hm+x, and then use the recursion relation Hn = Hn−1 + 1/n backwards m times, to unwind it to an approximation for Hx. Furthermore, this approximation is exact in the limit as m goes to infinity.


Specifically, for every integer n, we have that


limm→[Hm−Hm+n]=0,{displaystyle lim _{mrightarrow infty }left[H_{m}-H_{m+n}right]=0,,}{displaystyle lim _{mrightarrow infty }left[H_{m}-H_{m+n}right]=0,,}

and we can ask that the formula be obeyed if the arbitrary integer n is replaced by an arbitrary complex number x


limm→[Hm−Hm+x]=0.{displaystyle lim _{mrightarrow infty }left[H_{m}-H_{m+x}right]=0,.}{displaystyle lim _{mrightarrow infty }left[H_{m}-H_{m+x}right]=0,.}

Adding Hx to both sides gives


Hx=limm→[Hm−(Hm+x−Hx)]=limm→[(∑k=1m1k)−(∑k=1m1x+k)]=limm→k=1m(1k−1x+k)=x∑k=1∞1k(x+k).{displaystyle {begin{aligned}H_{x}&=lim _{mrightarrow infty }left[H_{m}-(H_{m+x}-H_{x})right]\[6pt]&=lim _{mrightarrow infty }left[left(sum _{k=1}^{m}{frac {1}{k}}right)-left(sum _{k=1}^{m}{frac {1}{x+k}}right)right]\[6pt]&=lim _{mrightarrow infty }sum _{k=1}^{m}left({frac {1}{k}}-{frac {1}{x+k}}right)=xsum _{k=1}^{infty }{frac {1}{k(x+k)}},.end{aligned}}}{displaystyle {begin{aligned}H_{x}&=lim _{mrightarrow infty }left[H_{m}-(H_{m+x}-H_{x})right]\[6pt]&=lim _{mrightarrow infty }left[left(sum _{k=1}^{m}{frac {1}{k}}right)-left(sum _{k=1}^{m}{frac {1}{x+k}}right)right]\[6pt]&=lim _{mrightarrow infty }sum _{k=1}^{m}left({frac {1}{k}}-{frac {1}{x+k}}right)=xsum _{k=1}^{infty }{frac {1}{k(x+k)}},.end{aligned}}}

This last expression for Hx is well defined for any complex number x except the negative integers, which fail because trying to use the recursion relation Hn = Hn−1 + 1/n backwards through the value n = 0 involves a division by zero. By construction, the function Hx is the unique function of x for which (1) H0 = 0, (2) Hx = Hx−1 + 1/x for all complex values x except the non-positive integers, and (3) limm→+∞ (Hm+xHm) = 0 for all complex values x.


Based on this last formula, it can be shown that:


01Hxdx=γ,{displaystyle int _{0}^{1}H_{x},dx=gamma ,,}{displaystyle int _{0}^{1}H_{x},dx=gamma ,,}

where γ is the Euler–Mascheroni constant or, more generally, for every n we have:


0nHxdx=nγ+ln⁡(n!).{displaystyle int _{0}^{n}H_{x},dx=ngamma +ln {(n!)},.}int _{0}^{n}H_{x},dx=ngamma +ln {(n!)},.


Special values for fractional arguments


There are the following special analytic values for fractional arguments between 0 and 1, given by the integral


=∫011−1−xdx.{displaystyle H_{alpha }=int _{0}^{1}{frac {1-x^{alpha }}{1-x}},dx,.}H_{alpha }=int _{0}^{1}{frac {1-x^{alpha }}{1-x}},dx,.

More values may be generated from the recurrence relation


=Hα1+1α,{displaystyle H_{alpha }=H_{alpha -1}+{frac {1}{alpha }},,}{displaystyle H_{alpha }=H_{alpha -1}+{frac {1}{alpha }},,}

or from the reflection relation


H1−αcot⁡α)−+11−α.{displaystyle H_{1-alpha }-H_{alpha }=pi cot {(pi alpha )}-{frac {1}{alpha }}+{frac {1}{1-alpha }},.}H_{1-alpha }-H_{alpha }=pi cot {(pi alpha )}-{frac {1}{alpha }}+{frac {1}{1-alpha }},.

For example:



H12=2−2ln⁡2{displaystyle H_{frac {1}{2}}=2-2ln {2}}H_{frac {1}{2}}=2-2ln {2}

H13=3−π23−32ln⁡3{displaystyle H_{frac {1}{3}}=3-{tfrac {pi }{2{sqrt {3}}}}-{tfrac {3}{2}}ln {3}}H_{frac {1}{3}}=3-{tfrac {pi }{2{sqrt {3}}}}-{tfrac {3}{2}}ln {3}

H23=32(1−ln⁡3)+3π6{displaystyle H_{frac {2}{3}}={tfrac {3}{2}}(1-ln {3})+{sqrt {3}}{tfrac {pi }{6}}}H_{frac {2}{3}}={tfrac {3}{2}}(1-ln {3})+{sqrt {3}}{tfrac {pi }{6}}

H14=4−π2−3ln⁡2{displaystyle H_{frac {1}{4}}=4-{tfrac {pi }{2}}-3ln {2}}H_{frac {1}{4}}=4-{tfrac {pi }{2}}-3ln {2}

H34=43−3ln⁡2+π2{displaystyle H_{frac {3}{4}}={tfrac {4}{3}}-3ln {2}+{tfrac {pi }{2}}}H_{frac {3}{4}}={tfrac {4}{3}}-3ln {2}+{tfrac {pi }{2}}

H16=6−π23−2ln⁡2−32ln⁡3{displaystyle H_{frac {1}{6}}=6-{tfrac {pi }{2}}{sqrt {3}}-2ln {2}-{tfrac {3}{2}}ln {3}}H_{frac {1}{6}}=6-{tfrac {pi }{2}}{sqrt {3}}-2ln {2}-{tfrac {3}{2}}ln {3}

H18=8−π2−4ln⁡2−12{π+ln⁡(2+2)−ln⁡(2−2)}{displaystyle H_{frac {1}{8}}=8-{tfrac {pi }{2}}-4ln {2}-{tfrac {1}{sqrt {2}}}left{pi +ln left(2+{sqrt {2}}right)-ln left(2-{sqrt {2}}right)right}}H_{frac {1}{8}}=8-{tfrac {pi }{2}}-4ln {2}-{tfrac {1}{sqrt {2}}}left{pi +ln left(2+{sqrt {2}}right)-ln left(2-{sqrt {2}}right)right}

H112=12−3(ln⁡2+ln⁡32)−π(1+32)+23ln⁡(2−3){displaystyle H_{frac {1}{12}}=12-3left(ln {2}+{tfrac {ln {3}}{2}}right)-pi left(1+{tfrac {sqrt {3}}{2}}right)+2{sqrt {3}}ln left({sqrt {2-{sqrt {3}}}}right)}H_{frac {1}{12}}=12-3left(ln {2}+{tfrac {ln {3}}{2}}right)-pi left(1+{tfrac {sqrt {3}}{2}}right)+2{sqrt {3}}ln left({sqrt {2-{sqrt {3}}}}right)


For positive integers p and q with p < q, we have:


Hpq=qp+2∑k=1⌊q−12⌋cos⁡(2πpkq)ln⁡(sin⁡kq))−π2cot⁡pq)−ln⁡(2q){displaystyle H_{frac {p}{q}}={frac {q}{p}}+2sum _{k=1}^{lfloor {frac {q-1}{2}}rfloor }cos left({frac {2pi pk}{q}}right)ln left({sin left({frac {pi k}{q}}right)}right)-{frac {pi }{2}}cot left({frac {pi p}{q}}right)-ln left(2qright)}{displaystyle H_{frac {p}{q}}={frac {q}{p}}+2sum _{k=1}^{lfloor {frac {q-1}{2}}rfloor }cos left({frac {2pi pk}{q}}right)ln left({sin left({frac {pi k}{q}}right)}right)-{frac {pi }{2}}cot left({frac {pi p}{q}}right)-ln left(2qright)}


Relation to the Riemann zeta function


Some derivatives of fractional harmonic numbers are given by:


dnHxdxn=(−1)n+1n![ζ(n+1)−Hx,n+1]dnHx,2dxn=(−1)n+1(n+1)![ζ(n+2)−Hx,n+2]dnHx,3dxn=(−1)n+112(n+2)![ζ(n+3)−Hx,n+3].{displaystyle {begin{aligned}{frac {d^{n}H_{x}}{dx^{n}}}&=(-1)^{n+1}n!left[zeta (n+1)-H_{x,n+1}right]\[6pt]{frac {d^{n}H_{x,2}}{dx^{n}}}&=(-1)^{n+1}(n+1)!left[zeta (n+2)-H_{x,n+2}right]\[6pt]{frac {d^{n}H_{x,3}}{dx^{n}}}&=(-1)^{n+1}{frac {1}{2}}(n+2)!left[zeta (n+3)-H_{x,n+3}right].end{aligned}}}{displaystyle {begin{aligned}{frac {d^{n}H_{x}}{dx^{n}}}&=(-1)^{n+1}n!left[zeta (n+1)-H_{x,n+1}right]\[6pt]{frac {d^{n}H_{x,2}}{dx^{n}}}&=(-1)^{n+1}(n+1)!left[zeta (n+2)-H_{x,n+2}right]\[6pt]{frac {d^{n}H_{x,3}}{dx^{n}}}&=(-1)^{n+1}{frac {1}{2}}(n+2)!left[zeta (n+3)-H_{x,n+3}right].end{aligned}}}

And using Maclaurin series, we have for x < 1:


Hx=∑n=1∞(−1)n+1xnζ(n+1)Hx,2=∑n=1∞(−1)n+1(n+1)xnζ(n+2)Hx,3=12∑n=1∞(−1)n+1(n+1)(n+2)xnζ(n+3).{displaystyle {begin{aligned}H_{x}&=sum _{n=1}^{infty }(-1)^{n+1}x^{n}zeta (n+1)\[5pt]H_{x,2}&=sum _{n=1}^{infty }(-1)^{n+1}(n+1)x^{n}zeta (n+2)\[5pt]H_{x,3}&={frac {1}{2}}sum _{n=1}^{infty }(-1)^{n+1}(n+1)(n+2)x^{n}zeta (n+3).end{aligned}}}{displaystyle {begin{aligned}H_{x}&=sum _{n=1}^{infty }(-1)^{n+1}x^{n}zeta (n+1)\[5pt]H_{x,2}&=sum _{n=1}^{infty }(-1)^{n+1}(n+1)x^{n}zeta (n+2)\[5pt]H_{x,3}&={frac {1}{2}}sum _{n=1}^{infty }(-1)^{n+1}(n+1)(n+2)x^{n}zeta (n+3).end{aligned}}}

For fractional arguments between 0 and 1, and for a > 1:


H1/a=1a(ζ(2)−1aζ(3)+1a2ζ(4)−1a3ζ(5)+⋯)H1/a,2=1a(2ζ(3)−3aζ(4)+4a2ζ(5)−5a3ζ(6)+⋯)H1/a,3=12a(2⋅(4)−3⋅4aζ(5)+4⋅5a2ζ(6)−5⋅6a3ζ(7)+⋯).{displaystyle {begin{aligned}H_{1/a}&={frac {1}{a}}left(zeta (2)-{frac {1}{a}}zeta (3)+{frac {1}{a^{2}}}zeta (4)-{frac {1}{a^{3}}}zeta (5)+cdots right)\[6pt]H_{1/a,,2}&={frac {1}{a}}left(2zeta (3)-{frac {3}{a}}zeta (4)+{frac {4}{a^{2}}}zeta (5)-{frac {5}{a^{3}}}zeta (6)+cdots right)\[6pt]H_{1/a,,3}&={frac {1}{2a}}left(2cdot 3zeta (4)-{frac {3cdot 4}{a}}zeta (5)+{frac {4cdot 5}{a^{2}}}zeta (6)-{frac {5cdot 6}{a^{3}}}zeta (7)+cdots right).end{aligned}}}{displaystyle {begin{aligned}H_{1/a}&={frac {1}{a}}left(zeta (2)-{frac {1}{a}}zeta (3)+{frac {1}{a^{2}}}zeta (4)-{frac {1}{a^{3}}}zeta (5)+cdots right)\[6pt]H_{1/a,,2}&={frac {1}{a}}left(2zeta (3)-{frac {3}{a}}zeta (4)+{frac {4}{a^{2}}}zeta (5)-{frac {5}{a^{3}}}zeta (6)+cdots right)\[6pt]H_{1/a,,3}&={frac {1}{2a}}left(2cdot 3zeta (4)-{frac {3cdot 4}{a}}zeta (5)+{frac {4cdot 5}{a^{2}}}zeta (6)-{frac {5cdot 6}{a^{3}}}zeta (7)+cdots right).end{aligned}}}


See also



  • Watterson estimator

  • Tajima's D

  • Coupon collector's problem

  • Jeep problem

  • Riemann zeta function

  • List of sums of reciprocals



Notes





  1. ^ ab
    John H., Conway; Richard K., Guy (1995). The book of numbers. Copernicus..mw-parser-output cite.citation{font-style:inherit}.mw-parser-output .citation q{quotes:"""""""'""'"}.mw-parser-output .citation .cs1-lock-free a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .citation .cs1-lock-limited a,.mw-parser-output .citation .cs1-lock-registration a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .citation .cs1-lock-subscription a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration{color:#555}.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration span{border-bottom:1px dotted;cursor:help}.mw-parser-output .cs1-ws-icon a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/4/4c/Wikisource-logo.svg/12px-Wikisource-logo.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output code.cs1-code{color:inherit;background:inherit;border:inherit;padding:inherit}.mw-parser-output .cs1-hidden-error{display:none;font-size:100%}.mw-parser-output .cs1-visible-error{font-size:100%}.mw-parser-output .cs1-maint{display:none;color:#33aa33;margin-left:0.3em}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-format{font-size:95%}.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-left{padding-left:0.2em}.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-right{padding-right:0.2em}



  2. ^ Ronald L., Graham; Donald E., Knuth; Oren, Patashnik (1994). Concrete Mathematics. Addison-Wesley.


  3. ^ Sondow, Jonathan and Weisstein, Eric W. "Harmonic Number." From MathWorld--A Wolfram Web Resource. http://mathworld.wolfram.com/HarmonicNumber.html


  4. ^ Sandifer, C. Edward (2007), How Euler Did It, MAA Spectrum, Mathematical Association of America, p. 206, ISBN 9780883855638.


  5. ^ Weisstein, Eric W. (2003). CRC Concise Encyclopedia of Mathematics. Boca Raton, FL: Chapman & Hall/CRC. p. 3115. ISBN 1-58488-347-2.


  6. ^ Eisenstein, Ferdinand Gotthold Max (1850). "Eine neue Gattung zahlentheoretischer Funktionen, welche von zwei Elementen ahhängen und durch gewisse lineare Funktional-Gleichungen definirt werden". Berichte Königl. Preuβ. Akad. Wiss. Berlin. 15: 36–42.


  7. ^ Eswarathasan, Arulappah; Levine, Eugene (1991). "p-integral harmonic sums". Discrete Mathematics. 91: 249–257. doi:10.1016/0012-365X(90)90234-9.


  8. ^ Boyd, David W. (1994). "A p-adic study of the partial sums of the harmonic series". Experimental Mathematics. 3: 287–302. doi:10.1080/10586458.1994.10504298.


  9. ^ Sanna, Carlo (2016). "On the p-adic valuation of harmonic numbers". Journal of Number Theory. 166: 41–46. doi:10.1016/j.jnt.2016.02.020.


  10. ^ Chen, Yong-Gao; Wu, Bing-Ling (2017). "On certain properties of harmonic numbers". Journal of Number Theory. 175: 66–86. doi:10.1016/j.jnt.2016.11.027.


  11. ^ Jeffrey Lagarias (2002). "An Elementary Problem Equivalent to the Riemann Hypothesis". Amer. Math. Monthly. 109: 534–543. arXiv:math.NT/0008177. doi:10.2307/2695443.




References




  • Arthur T. Benjamin; Gregory O. Preston; Jennifer J. Quinn (2002). "A Stirling Encounter with Harmonic Numbers" (PDF). Mathematics Magazine. 75 (2): 95–103. doi:10.2307/3219141.


  • Donald Knuth (1997). "Section 1.2.7: Harmonic Numbers". The Art of Computer Programming. Volume 1: Fundamental Algorithms (Third ed.). Addison-Wesley. pp. 75–79. ISBN 0-201-89683-4.

  • Ed Sandifer, How Euler Did It — Estimating the Basel problem (2003)


  • Paule, Peter; Schneider, Carsten (2003). "Computer Proofs of a New Family of Harmonic Number Identities" (PDF). Adv. Appl. Math. 31 (2): 359–378. doi:10.1016/s0196-8858(03)00016-2.


  • Wenchang Chu (2004). "A Binomial Coefficient Identity Associated with Beukers' Conjecture on Apery Numbers" (PDF). The Electronic Journal of Combinatorics. 11: N15.


  • Ayhan Dil; István Mező (2008). "A Symmetric Algorithm for Hyperharmonic and Fibonacci Numbers". Applied Mathematics and Computation. 206 (2): 942–951. arXiv:0803.4388. doi:10.1016/j.amc.2008.10.013.



External links


  • Weisstein, Eric W. "Harmonic Number". MathWorld.

This article incorporates material from Harmonic number on PlanetMath, which is licensed under the Creative Commons Attribution/Share-Alike License.







這個網誌中的熱門文章

12.7 cm/40 Type 89 naval gun

Shark

Wiciokrzew